Biological Conservation 181 (2015) 111–123

Contents lists available at ScienceDirect

Biological Conservation

journal homepage: www.elsevier.com/locate/biocon

Review The ecological significance of giant in reef ⇑ Mei Lin Neo a,b, William Eckman a, Kareen Vicentuan a,b, Serena L.-M. Teo b, Peter A. Todd a, a Experimental Marine Ecology Laboratory, Department of Biological Sciences, National University of , 14 Science Drive 4, Singapore 117543, Singapore b Tropical Marine Science Institute, National University of Singapore, 18 Kent Ridge Road, Singapore 119227, Singapore article info abstract

Article history: Giant clams ( and ) are thought to play various ecological roles in Received 14 May 2014 ecosystems, but most of these have not previously been quantified. Using data from the literature and Received in revised form 29 October 2014 our own studies we elucidate the ecological functions of giant clams. We show how their tissues are food Accepted 2 November 2014 for a wide array of predators and scavengers, while their discharges of live , faeces, and Available online 5 December 2014 gametes are eaten by opportunistic feeders. The shells of giant clams provide substrate for colonization by epibionts, while commensal and ectoparasitic organisms live within their cavities. Giant clams Keywords: increase the topographic heterogeneity of the reef, act as reservoirs of zooxanthellae ( spp.), Carbonate budgets and also potentially counteract eutrophication via water filtering. Finally, dense populations of giant Conservation Epibiota clams produce large quantities of shell material that are eventually incorporated into Eutrophication the reef framework. Unfortunately, giant clams are under great pressure from overfishing and extirpa- Giant clams tions are likely to be detrimental to coral reefs. A greater understanding of the numerous contributions Hippopus giant clams provide will reinforce the case for their conservation. Tridacna Ó 2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license Zooxanthellae (http://creativecommons.org/licenses/by-nc-nd/3.0/).

Contents

1. Introduction ...... 112 2. Methods ...... 112 2.1. Literature survey ...... 112 2.2. Population estimates of ecologically relevant parameters ...... 112 3. Giant clams as food ...... 114 3.1. Productivity and biomass ...... 114 3.2. Food for predators and scavengers...... 115 3.3. Expelled materials...... 115 4. Giant clams as shelter ...... 117 4.1. Shelters for coral reef fish...... 117 4.2. Shell surfaces for epibionts...... 117 4.3. Hosts for ectoparasites ...... 118 4.4. Hosts for commensals ...... 118 5. Reef-scale contributions of giant clams ...... 118 5.1. Contributors of carbonate ...... 118 5.2. Bioeroders ...... 118 5.3. Topographic enhancement ...... 118 5.4. Source of zooxanthellae (Symbiodinium spp.) ...... 119 5.5. Counteractors of eutrophication...... 120 6. Conclusion ...... 120

⇑ Corresponding author. Tel.: +65 65161034. E-mail addresses: [email protected] (M.L. Neo), [email protected] (P.A. Todd). http://dx.doi.org/10.1016/j.biocon.2014.11.004 0006-3207/Ó 2014 The Authors. Published by Elsevier Ltd. This is an open access article under the CC BY-NC-ND license (http://creativecommons.org/licenses/by-nc-nd/3.0/). 112 M.L. Neo et al. / Biological Conservation 181 (2015) 111–123

Acknowledgements ...... 121 Appendix A. Supplementary material ...... 121 References ...... 121

1. Introduction however, stocks are declining rapidly in many countries (bin Othman et al., 2010; Andréfouët et al., 2013) and extirpations are As recently summarized by Bridge et al. (2013, p.528) coral occurring (Kinch and Teitelbaum, 2010; Neo and Todd, 2012, reefs globally are ‘‘suffering death by a thousand cuts’’. Some of 2013). these, including global warming and acidification, are There exists a substantial body of work on the and notorious and possibly fatal. Others, such as the loss of particular of giant clams, but their significance in the coral reef species or genera, are generally less pernicious and do not garner is not well understood. Some previous researchers have the same attention. Of course, all reef organisms have a role to play provided anecdotal insights into their likely roles, i.e. as food, as but giant clams (Cardiidae: ), by virtue of their sheer shelter, and as reef-builders and shapers. For example, Mercier size (Yonge, 1975), well developed with zooxanthellae and Hamel (1996, p.113) remarked: ‘‘Tridacna face many dangers. (Yonge, 1980), and highly threatened status throughout much of They are most vulnerable early in their life cycle, when they are prey their geographic range (Lucas, 1994), perhaps deserve special con- to , lobsters, wrasses, pufferfish, and eagle rays.’’ In a popular sideration. Based on tridacnine taxa, these iconic inverte- science article, Mingoa-Licuanan and Gomez (2002, p.24) com- brates have been associated with since the late Eocene mented: ‘‘ populations add topographic detail to the seabed (Harzhauser et al., 2008) and facies of more recent Tridacna species and serve as nurseries to various organisms... Their calcified shells are common in the upper strata of fossilized reefs (Accordi et al., are excellent substrata for sedentary organisms.’’ Finally, Hutchings 2010; Ono and Clark, 2012). Modern giant clams are only found (1986, p.245) stated: ‘‘giant clams are recognisable in early in the Indo-West Pacific (Harzhauser et al., 2008) in the area reefs and if similar densities occurred to those on recent reefs, giant bounded by southern Africa, the Red Sea, , Polynesia (exclud- clams have had a considerable ongoing impact on reef .’’ ing New Zealand and ), and (bin Othman et al., Even though there is evidence that giant clams contribute to the 2010). There are currently 13 extant species of giant clams (see functioning of coral reefs, this has very rarely been quantified. Table 1), including two recently rediscovered: Tridacna noae (Su Cabaitan et al. (2008) represents the only study to experimentally et al., 2014; Borsa et al., 2014) and Tridacna squamosina (previously demonstrate the benefits that giant clams can have on coral reefs. known as T. costata)(Richter et al., 2008), one new species: They showed that, compared to control plots, the presence of clams Tridacna ningaloo (Penny and Willan, 2014), and an undescribed had significantly positive effects on the richness and abundance of cryptic Tridacna sp. (Huelsken et al., 2013). Tridacna maxima is fish species and various . Here, based on existing lit- the most widespread, while , Tridacna mba- erature and our own observations, we examine giant clams as con- lavuana (previously known as T. tevoroa), T. ningaloo, T. noae, Trid- tributors to reef productivity, as providers of biomass to predators acna rosewateri, and T. squamosina have much more restricted and scavengers, and as nurseries and hosts for other organisms. We distributions (Rosewater, 1965; bin Othman et al., 2010; Penny also examine their reef-scale roles as calcium carbonate producers, and Willan, 2014; Su et al., 2014). Tridacna gigas is by far the largest zooxanthellae reservoirs, and counteractors of eutrophication. Our species, reaching shell lengths of over 120 cm and weights in findings lead to the conclusion that healthy populations of giant excess of 200 kg (Rosewater, 1965). Since pre-history, giant clams’ clams benefit coral reefs in ways previously underappreciated, high biomass and heavy calcified shells have made them useful to and that this knowledge should help prioritize their conservation. humans as a source of food and material (Miller, 1979; Hviding, 1993). However, as a result of degradation, technological 2. Methods advances in exploitation, expanding trade networks, and demand by aquarists, numbers are declining throughout their 2.1. Literature survey range (Mingoa-Licuanan and Gomez, 2002; Kinch and Teitelbaum, 2010; bin Othman et al., 2010). In this review, we first drew upon our own archives of publica- Giant clams are especially vulnerable to stock depletion tions, proceedings, dissertations, books, manuals, technical reports, because of their late sexual maturity, sessile adult phase, and popular science magazines, and grey literature that have been col- broadcast spawning strategy (Munro, 1989; Lucas, 1994). Fertiliza- lected during more than 10 years of giant clam research. These tion success requires sufficient numbers of spawning individuals, archives (n = 481 publications) were supplemented with key-word and low densities result in reduced (or zero) recruitment and even- searches in five major literature databases, i.e. Google Scholar, tual population collapse (Braley, 1984, 1987; Neo et al., 2013). JSTOR, PubMed, ScienceDirect, and Web-of-Science. We also used Presently, all giant clam species, other than the new species, T. ‘‘snowball’’ sampling (see Lescureux and Linnell, 2014), that is, ningaloo, the recently rediscovered T. noae and T. squamosina, and we manually searched through the reference lists of the most rel- the cryptic Tridacna sp., are protected under Appendix II of the evant giant clam papers to identify (and subsequently retrieve) Convention on International Trade in Endangered Species of Wild some of the more obscure literature. and Flora (CITES) and listed in the IUCN Red List of Threa- tened Species (Table 1). Conservation efforts are ongoing 2.2. Population estimates of ecologically relevant parameters (Heslinga, 2013) including essential basic research (e.g. restocking of clams in heavily impacted coral reefs, Guest et al., 2008; effects For all the estimates described below, we first identified surveys of shade on survival and growth of juvenile clams, Adams et al., of natural giant clams that included both population density and 2013; early chemotaxis contributing to active habitat selection, size distribution (i.e. Pearson and Munro, 1991; Chantrapornsyl Dumas et al., 2014) and the development of new restocking et al., 1996; Black et al., 2011; Gilbert et al., 2006; Todd et al., techniques (Waters et al., 2013). There are also several giant clam 2009). All densities were converted to per hectare values prior to sanctuaries under legal protection, for example, in Australia (Rees the calculations. The reported size distributions did not provide et al., 2003) and (Andréfouët et al., 2005, 2013); individual measurements for each clam; rather they stated the Table 1 Giant clam species list (Rosewater, 1965; Richter et al., 2008; bin Othman et al., 2010; Borsa et al., 2014; Huelsken et al., 2013; Penny and Willan, 2014; Su et al., 2014) and their categories listed by the International Union for Conservation of Nature (IUCN) Red List of Threatened Species (Molluscs Specialist Group, 1996; Wells, 1996).

Species name Description Global conservation status (Linnaeus, 1758) Species has strong radial ribbing and reddish blotches in irregular bands on shells, growing to about 40 cm. Unlike Lower risk/conservation dependent Tridacna species, Hippopus mantle does not extend over shell margins and has a narrow byssal orifice 111–123 (2015) 181 Conservation Biological / al. et Neo M.L. Hippopus porcellanus Rosewater, 1982 Species is distinguished from H. hippopus by its smoother and thinner shells, and presence of fringing at Lower risk/conservation dependent incurrent , growing to approximately 40 cm Lamarck, 1819 Smallest of all clam species, reaching lengths of about 15 cm. Burrows and completely embeds into reef substrates Lower risk/least concern (Röding, 1798) Second largest species, growing up to 60 cm. Has heavy and plain shells, with no strong ribbing Vulnerable A2cd Tridacna gigas (Linnaeus, 1758) Largest of all clam species, growing to over 1 m long. Easily identified by their size and elongate, triangular projections Vulnerable A2cd of upper shell margins Tridacna maxima (Röding, 1798) Species is identified by its close-set scutes. Grows up to 35 cm. Tends to bore partially into reef substrates Lower Risk/conservation dependent Tridacna mbalavuana Ladd, 1934 Species is most like T. derasa in appearance, but distinguished by its rugose mantle, prominent guard tentacles present Vulnerable B1 + 2c (formerly T. tevoroa Lucas, Ledua, on the incurrent siphon, thinner valves, and coloured patches on shell ribbing. Can grow over 50 cm long. Restricted to Braley, 1990) Fiji and Tonga Tridacna rosewateri Sirenko and Species is most like T. squamosa in appearance, but distinguished by its thinner shell, large byssal orifice and dense Vulnerable A2cd Scarlato, 1991 scutes on primary radial folds. Only found in Mauritius, with largest specimen measured at 19.1 cm Lamarck, 1819 Species is identified by its large, well-spaced scutes, with shell lengths up to 40 cm Lower risk/conservation dependent Tridacna ningaloo Penny and Willan, Species is most like T. maxima in appearance; weakly differentiated morphologically but strongly defined genetically. Not assessed 2014 n. sp. Holotype specimen measures 17.9 cm. Current known distribution in Western Australia, but possibly extends to Tridacna noae (Röding, 1798) Species is most like T. maxima in appearance, but distinguished by its sparsely distributed hyaline organs and oval Not assessed patches with different colors bounded by white margins along mantle edge. Shell lengths between 6 and 20 cm. Overlapping distributions with T. maxima but generally in lower abundances Tridacna squamosina Sturany, 1899 Species is most like T. squamosa in appearance, but distinguished by its crowded, well-spaced scutes, asymmetrical Not assessed (formerly T. costata Roa-Quiaoit, shell, and grows up to 32 cm. Only found in the Red Sea Kochzius, Jantzen, Zibdah, Richter, 2008) Cryptic Tridacna sp. (undescribed in Recently determined as a widely distributed cryptic species; forms an evolutionarily distinct monophyletic group Not assessed Huelsken et al., 2013) 113 114 M.L. Neo et al. / Biological Conservation 181 (2015) 111–123

number of clams in a series of size brackets or ‘bins’. For the of 0.145 were used (Pearson and Munro, 1991). For T. maxima, L1 purpose of our analysis, we assumed the size of each clam in a of 27.8, K of 0.068, and t0 of 0 were used (Black et al., 2011). Once bin was equal to (minimum bin size + maximum bin size)/2. For the ages of the clams were estimated, one year was added, and the the following equations, mass and weight are in grams, and shell von Bertalanffy equation was used to predict a new shell length. length is abbreviated to SL. The previous calculations were then used to predict biomass and Standing tissue biomass calculations: for each size T. maxima, shell weight of the clams at these increased sizes, and annual bio- wet biomass was first determined using the formulae biomass = mass and shell production were assumed to be the differences

10^[(3.0434 LOG(SLcm))1.6026] (Gilbert et al., 2006) for clams between the estimated values on the survey date and the predicted at Tatakoto atoll and biomass = 10^[(2.9367 LOG(SLcm))1.5318] values one year in the future. for clams at Fangatau atoll and Ningaloo reefs (Gilbert et al., 2006) Clearance rate (CR) calculations: for T. gigas, clearance rate for a and then assumed that 5.8% of wet biomass would convert to dry single clam (l h1) was calculated using the formula CR = 3.68 biomass (Ricciardi and Bourget, 1998). For each size T. crocea, dry (dry weight0.397), while for T. crocea the formula used was biomass was determined using the formula biomass = CR = 0.585 (dry weight0.905)(Klumpp and Griffiths, 1994). No 6 3.24 (3.23 10 ) (SLmm)(Klumpp and Griffiths, 1994), while for T. formula was available for T. maxima. The clearance rate for each 6 3.36 gigas, the formula biomass = (0.34 10 ) (SLmm) was used size of clams was then multiplied by the number of clams (Klumpp and Griffiths, 1994). The biomass for each size class of of that size, and the multiplied values were totalled. clams was then multiplied by the number of clams of that size, and the multiplied values were totalled. Standing shell weight calculations: for each size T. maxima, total 3. Giant clams as food clam weight was determined using the formulae weight =

10^[(3.1335 LOG(SLcm)) 0.9173] for clams at Tatakoto atoll 3.1. Productivity and biomass (Gilbert et al., 2006) and weight = 10^[(3.1634 LOG(SLcm)] 0.9495) for clams at Fangatau atoll and Ningaloo reefs (Gilbert Giant clams are mixotrophic (Jantzen et al., 2008), being capa- et al., 2006) and then subtracting wet tissue weight (calculated ble of generating biomass through both primary and secondary above) to give the weight of the shell alone. For each size T. crocea, production. Primary production is controlled by the photosynthetic shell weight was determined using the formula weight = efficiency of their symbiotic photoautotrophic zooxanthellae 5 3.51 (2.05 10 ) (SLmm )(Klumpp and Griffiths, 1994), while for (Jantzen et al., 2008; Yau and Fan, 2012). Secondary production, 5 3.11 T. gigas, the formula weight = (4.76 10 ) (SLmm) was used on the other hand, is strongly influenced by the uptake rate of (Klumpp and Griffiths, 1994). The shell weight for each size class ambient dissolved inorganic carbon (DIC) via filter feeding (Jones of clams was then multiplied by the number of clams of that size, et al., 1986; Watanabe et al., 2004). The acquisition of DIC is related and the multiplied values were totalled. to clearance rates (i.e. the volume of water each clam pumps per Annual biomass and shell production calculations: to determine unit time), and therefore clam body size (Klumpp et al., 1992). To annual biomass and shell production, von Bertalanffy equations facilitate between-taxa comparisons, the net primary productivity (available for T. gigas and T. maxima, but not T. crocea) were used (NPP) from an array of reef organisms, including giant clams, is to estimate the age of each size clam from four locations presented in Fig. 1. We acknowledge that different productivity (Pearson and Munro, 1991; Black et al., 2011; Gilbert et al., measures were used across studies; however, our aim is to provide 2006). For T. gigas, growth parameter estimates: asymptotic length estimate figures for relative rates among reef organisms. The NPP 2 1 (L1) of 80, growth (K) of 0.105, and its theoretical date of ‘birth’ (t0) of the giant clams, T. maxima (28.16 g O2 m d ) and T. squamosa

2 1 Fig. 1. Maximum net primary productivity (NPP) of different reef flora and fauna, measured in terms of net oxygen production (units = g O2 m d ). NPP values are arranged from the highest to lowest producers. Standard deviation provided when available. Information extracted from: Wanders (1976), Rogers and Salesky (1981), Porter et al. (1984), Chisholm (2003), Jantzen et al. (2008), Naumann et al. (2013). M.L. Neo et al. / Biological Conservation 181 (2015) 111–123 115

Table 2 Estimates of ecologically relevant parameters of giant clam populations found per hectare of reef area (based on data extracted from the references cited in the table). DD = data deficient.

Location Population Standing Annual biomass Shell Annual shell Water Source of density biomass production weight (kg) production (kg) filtration (Lh1) population data (individuals) (kg dry (kg dry weight) weight) Tridacna crocea Lee-Pae Island, Andaman Sea, 2,441 17 DD 391 DD 8,144 Chantrapornsyl et al. (1996) Tioman Island, 955 4 DD 98 DD 2,115 Todd et al. (2009) Tridacna maxima Fangatau atoll, French Polynesia 381,919 878 217 89,023 23,372 DD Gilbert et al. (2006) Tatakoto atoll, French Polynesia 909,466 1,041 238 102,833 37,040 DD Gilbert et al. (2006) Ningaloo Marine Park, Western Australia 8,600 36 7 3,898 562 DD Black et al. (2011) Tridacna gigas , Australia 432 718 14 18,839 356 28,121 Pearson and Munro (1991)

2 1 (18.14 g O2 m d ) is greater than most of the other coral reef apparent that giant clams are widely utilized food sources on coral primary producers. From the examples in Fig. 1, the NPP of T. max- reefs, with 75 known predators (Table 3). Fishes—wrasse, trigger- ima and T. squamosa are respectively 74.1 and 47.7 higher fish, and pufferfish—prey on both juvenile and adult giant clams than the lowest NPP presented—that of the hard coral (Manicina (Alcazar, 1986; Richardson, 1991; Govan, 1992b), and bite marks 2 1 sp.) (0.38 g O2 m d )—and approximately double that of the on the mantle edges of wild clams are common (Fig. 2). In maricul- relatively fast growing branching coral Acropora palmata. The ture, ectoparasitic pyramidellids and ranellids are often abundant contribution of giant clams to overall reef productivity is hence and their attacks can devastate juvenile cohorts (Perron et al., potentially very substantial, especially when populations are dense 1985; Boglio and Lucas, 1997), but they have less impact on clams (Rees et al., 2003; Andréfouët et al., 2005; Gilbert et al., 2006). on reefs, where natural predators of these ectoparasites are present It is known that cultured stocks can produce substantial bio- (Cumming and Alford, 1994; Govan, 1995). mass, e.g. 29 t ha1 yr1 of wet tissue biomass for T. gigas (Barker The wide array of morphological and behavioural defences et al., 1988) (estimated 1,682 kg dry weight ha1 yr1) and exhibited by giant clams (Soo and Todd, 2014) is also indicative 16 t ha1 yr1 of wet tissue biomass for T. derasa (Heslinga et al., of their importance as a food source. Giant clams and their 1984) (estimated 928 kg dry weight ha1 yr1). Here, we provide predators are likely to have been in an evolutionary arms race estimates of tissue biomass for natural populations of three giant for millions of years. To resist attack, tridacnines have evolved clam species (Table 2). We also estimated annual biomass produc- large body sizes (Carter, 1968), reduced byssal orifices, and heavy tion which, if the giant clam populations were in equilibrium, strong shells (Perron et al., 1985; Alcazar, 1986; Govan et al., would equal the amount of food provided to predators and scav- 1993). Neo and Todd (2011a) found that shell strength is a pheno- engers per year. Giant clams will contribute more to productivity typically plastic trait in juvenile T. squamosa, with specimens on reefs where there is recruitment of juvenile clams, as these exposed to predator effluents being harder to crush. The shell are faster-growing. In French Polynesia, the Tatakoto atoll popula- projections (called scutes) in some Tridacna species probably offer tion of T. maxima, a medium-sized species, has a high standing crop protection from crushing predators such as crabs and jawed fishes (1,041 kg dry weight ha1) and high productivity, being capable of (Ling et al., 2008). Other defence mechanisms include aggregation producing 238 kg dry weight ha1 yr1 of biomass. This population of conspecifics (Huang et al., 2007), camouflage (Todd et al., 2009), is maintained by especially rapid recruitment, probably due to rapid mantle withdrawal (McMichael, 1974), and squirting of thermal variations caused by the geography of the atoll (Gilbert water from (Neo and Todd, 2011b). et al., 2006). The example T. gigas population from the Great Barrier The scavenging guild is critical to nutrient recycling on coral Reef (Table 2) has a standing crop of 718 kg dry weight ha1, but is reefs (Keable, 1995; Rassweiler and Rassweiler, 2011) and dead essentially a relict population, consisting primarily of large adult or dying giant clams will attract a variety of small clams. The lack of younger, faster-growing T. gigas clams explains scavengers including isopods, , amphipods, leptostracans, why the annual production of new biomass is so low (14 kg dry mysids, polychaetes, and small decapods and (Keable, 1995). weight ha1 yr1). Tridacna crocea appears to contribute minimally Many of these have not been reported to prey on healthy clams; for on a per hectare basis (due to its smaller size and low population example, the muricid gastropod (Drupella rugosa) only acts as a density) in the examples provided in Table 2, but in patches of scavenger on giant clam juveniles (Perron et al., 1985). favourable habitat, T. crocea can have densities exceeding 100 clams m2 (Hamner and Jones, 1976) and hence may be 3.3. Expelled materials important at very local scales. While we have only presented data for single species, it is possible for up to six to co-exist on the same Opportunistic feeders may consume the materials (gametes, reef (e.g. Hardy and Hardy, 1969; Rees et al., 2003), occupying dif- faeces, and pseudofaeces) expelled by giant clams (Ricard and ferent niches based on depth and substrate type. Salvat, 1977; Lucas, 1994). For example, at the Silaqui ocean nurs- ery, Bolinao, , a large school of blue sprat (Spratelloides 3.2. Food for predators and scavengers delicatulus) fed for at least three hours on the gametes released by T. gigas (Maboloc and Mingoa-Licuanan, 2011). Routine releases on juvenile giant clams has been studied extensively of undigested, photosynthetically functional zooxanthellae in the (e.g. Alcazar, 1986; Perio and Belda, 1989; Govan et al., 1993), faeces (Ricard and Salvat, 1977; Trench et al., 1981) can be impor- particularly during the ocean nursery phase of mariculture tant sources of organic matter in closed or semi-closed systems, (Govan, 1992a). Heslinga and Fitt (1987) noted that adult tridac- such as the atoll in French Polynesia (Ricard and Salvat, nines appeared to be, more-or-less, immune to predation, but there 1977; Richard, 1977). Finally, giant clams faeces contain substan- have been reported attacks on mature adults (Alcazar, 1986). It is tial amounts of nutritious and protein (Ricard and Salvat, 116 M.L. Neo et al. / Biological Conservation 181 (2015) 111–123

Table 3 Predators of giant clams, including those listed by Govan (1992a,b), plus new observations and additional findings from grey literature.

Predator species Method of predation Literature source(s) PORIFERA: Clionaidae (Boring ) Unknown Bore into shells, weakening shells Govan (1992b) : Family Turbellaria Stylochus (Imogene) matatasi Enter the clam through either the byssal orifice or inhalant siphon Newman et al. (1991,1993) Stylochus (Imogene) sp. Govan (1992a,b) Polyclad sp. 1 Govan (1992a) MOLLUSCS: Family () Cantharus fumosus – Perio and Belda (1989) Family Costellariidae (Mitres) Vexillum cruentatum – Govan (1992b) V. plicarium – Richardson (1991) Family (Tulip snails) Pleuroploca trapezium Immobilize clam by clasping mantle with foot preventing Govan (1992b) Pleuroploca sp. closure, insert proboscis into soft tissues Alcazar (1986) Family (Murexes) Chicoreus brunneus Drill holes into shells of juvenile clams Abdon-Naguit and Alcazar (1989), Govan (1992a,b) C. microphyllum Drill holes into shells Govan (1992a,b) C. palmarosae Often drill through valves; may attack via valve gape or byssal orifice Govan et al. (1993) C. ramosus Insert proboscis into byssal gape to reach soft tissues, inject paralytic Heslinga et al. (1984), Alcazar (1986), Govan (1992b) substance Cronia fiscella Drill holes into shells of juvenile clams Govan (1992b) C. margariticola Through valve gape Govan (1992a,b) C. ochrostoma Drill holes into shells Govan (1992b) Morula granulata Drill holes into shells Govan (1992a,b) Muricodrupa fiscella Drill holes into shells Govan (1992a) Thais aculeata Attack through valve gape Govan (1992a,b) Family Octopodidae () Octopus sp. Chip shells; pry valves apart to feed Heslinga et al. (1984), Barker et al. (1988), Govan (1992b), Mercier and Hamel (1996) Family Turbonilla sp. Use their long, flexible proboscis to suck clams’ body fluids, either Govan (1992a,b) Tathrella iredalei from mantle edge or through byssal orifice Heslinga et al. (1990), Govan (1992b) Family Ranellidae (Tritons) Bursa granularis Insert proboscis between valves of prey Govan et al. (1993) Cymatium aquatile Injection of an immobilizing fluid through mantle or byssal orifice, Abdon-Naguit and Alcazar (1989), Govan then feed on soft tissues (1992a,b,1995) C. muricinum Perron et al. (1985), Govan (1992a,b,1995) C. nicobaricum Govan (1992a,b,1995) C. pileare Govan (1992a,b,1995) C. vespaceum Perio and Belda (1989), Govan (1992b) Family (Volutes) amphora – Loch (1991) Melo sp. – Govan (1992b) Seastar Exert powerful suction and tire adductor muscles (pry open clam) Weingarten (1991) : Family Diogenidae (Hermit crabs) Dardanus deformis Crushed 26 juvenile T. gigas in 3 days Heslinga et al. (1984) D. lagopodes Chip valve ends Govan (1992a) D. pedunculatus Crush or chip valves of prey Govan (1992a), Govan et al. (1993) Family Gonodactylidae (Mantis ) Gonodactylus chiragra Smash shells Govan (1992a) Gonodactylus sp. – Govan (1992b) Family Portunidae (Swimming crabs) Thalamita admete Chip shells; attack via byssal orifice Govan (1992a) T. coerulipes Govan (1992a) T. crenata Crush shells; may pry clam open via ventral margin Ling (2007) T. danae Crush or chip valves; attack via byssal orifice of clams Govan et al. (1993) T. spinimana – Richardson (1991) T. stephensoni Chip shells; attack via byssal orifice Govan (1992a) T. cf. tenuipes Govan (1992a) Thalamita sp. Penetrate soft tissues of adults through either byssal orifice or the Alcazar (1986), Govan (1992b) inhalant siphon Family (Stone crabs) Atergatis floridus Crush or chip valves Richardson (1991), Govan et al. (1993) A. integerrimus Richardson (1991) Atergatis spp. Govan (1992b) Carpilius convexus Crush or chip valves of juvenile clams Alcazar (1986), Govan (1992a,b), Govan et al. (1993) C. maculatus Crush shells Govan (1992a,b) M.L. Neo et al. / Biological Conservation 181 (2015) 111–123 117

Table 3 (continued)

Predator species Method of predation Literature source(s) Demania cultripes Crush shells of juvenile clams Alcazar (1986), Govan (1992b) Leptodius sanguineus Crush shells Govan (1992a,b) pictor Crush or chip shells Richardson (1991), Govan (1992b) Myomenippe hardwickii Crush shells; may attack via byssal orifice Ling (2007) Zosimus aeneus Crush shells Govan (1992a,b) FISH: Family Balistidae (Triggerfish) Balistapus undulatus Feed on mantle and the exposed and foot of adult clams Alcazar (1986), Perio and Belda (1989) Balistoides viridescens Crush or chip shells Heslinga et al. (1990) Balistoides sp. Govan (1992b) Pseudobalistes flavimarginatus Heslinga et al. (1990), Chambers (2007) Pseudobalistes sp. Govan (1992b) Rhinecanthus sp. Govan (1992b) Family Lethrinidae (Emperors) Monotaxis grandoculis Directly consumed 50 juvenile T. squamosa in <2 h Heslinga et al. (1984), Govan (1992b) Family Labridae (Wrasses) Cheilinus fasciatus – Richardson (1991) Cheilinus sp. Crush or chip shells Govan (1992b) Choerodon anchorago – Richardson (1991) C. schoenleinii – Richardson (1991) Choerodon sp. Crush or chip shells Govan (1992b) Halichoeres sp. Feed only on the byssus and foot of unanchored clams Alcazar (1986), Govan (1992b) Thalassoma hardwicke – Richardson (1991) T. lunare – Richardson (1991) Family Myliobatidae (Eagle rays) Aetobatis narinari Crush shells Heslinga et al. (1990), Govan (1992b), Chambers (2007) Family Tetraodontidae (Pufferfish) Canthigaster solandri – Richardson (1991) C. valentini Crush or chip shells Perio and Belda (1989), Govan (1992b) Tetraodon stellatus Heslinga et al. (1990), Govan (1992b), Chambers (2007) TURTLES: Family Cheloniidae Caretta caretta – Bustard (1972) Chelonia mydas Break off shell flukes and ingest as calcium carbonate dietary Weingarten (1991) supplement

1977) that can make a significant dietary contribution to reef fish; be common in marine ecosystems (Harder, 2008), only a handful of the adult black damselfish (), for instance studies have discussed its ecological importance (e.g. Abellö et al., (Chan, 2007). 1990; Creed, 2000; Botton, 2009). Giant clam shells have been reported to harbour a variety of burrowing (Yonge, 1955; Turner 4. Giant clams as shelter and Boss, 1962) and encrusting (Roscoe, 1962; Rosewater, 1965) reef inhabitants, although the authors of these studies did not list 4.1. Shelters for coral reef fish specific taxa. Our own observations of clam-associated epibionts include macroalgae, sponges, ascidians, nudibranchs, bryozoans, Coral reef fish diversity is related to coral cover (Bell and Galzin, tubeworms, hard and soft corals, as well as small mobile inverte- 1984; Ault and Johnson, 1998) and substrate complexity brates (Fig. 3). Some, such as macroalgae (Fatherree, 2006), boring (Gratwicke and Speight, 2005; Lingo and Szedlmayer, 2006). Dense sponges (Norton et al., 1993), the boring worm (Oenone fulgida) aggregations of giant clams can increase topographic heterogene- (Delbeek and Sprung, 1994), and pest anemones (Aiptasia spp.) ity of the seabed and serve as nurseries and shelters for fishes. A (Fatherree, 2006) can harm their tridacnine hosts. In addition, foul- restoration study in the Philippines demonstrated that T. gigas ing on juvenile clams can reduce growth and lead to death by introduced onto degraded reefs significantly improved fish diver- interfering with valve movement (unpublished data). Conversely, sity and abundance compared to control plots (Cabaitan et al., other epibionts may protect their hosts by contributing anti-pred- 2008). An increase in habitat relief usually facilitates recruitment ator defenses (Feiferak, 1987) and/or camouflage (Harder, 2008). and settlement of juvenile fish and helps reduce predation by pro- Vicentuan-Cabaitan et al. (2014) identified the community liv- viding refuges (Beukers and Jones, 1997; Lecchini et al., 2007) ing on the valves of T. squamosa in Singapore. They found at least while the shell ridges of giant clams represent suitable obscure 49 species belonging to a minimum of 36 families living on the surfaces for the deposition of fishes’ masses (Weingarten, shells of eight T. squamosa individuals (shell lengths 236– 1991). The large mantle cavities of tridacnines also afford shelter 400 mm). Vicentuan-Cabaitan et al. (2014) also highlighted that to smaller fishes, such as the pearlfish (Encheliophis homei)(Trott giant clam shells provide much more surface area for colonization and Chan, 1972) and anemone fishes in the absence of compared to the patch of substrate they occupy (a 26:1 ratio based anemones (Arvedlund and Takemura, 2005). on three adult T. squamosa specimens). A complete taxa checklist was not provided in their short paper, but it is now included here 4.2. Shell surfaces for epibionts (Table A1). As this list is for just one tridacnine species at a single locality, we expect that it represents only a small percentage of the On coral reefs, where settlement surfaces are limiting, epibiosis taxa that live on the shells of all giant clam species (that vary in is an alternative colonization strategy for sessile organisms (Wahl size, shell morphology, habitat preference, and global distribution) and Mark, 1999; Harder, 2008). Nevertheless, while epibiosis may (see Fig. 3). 118 M.L. Neo et al. / Biological Conservation 181 (2015) 111–123

to T. gigas (Bruce, 1983, 2000). Due to their long lifespans, giant clams can host many generations of commensals and the absence of any trauma to collected examples suggests that life within tridacnines is secure (Bruce, 2000).

5. Reef-scale contributions of giant clams

5.1. Contributors of carbonate

The calcium carbonate framework of coral reefs is maintained by opposing processes of carbonate production and removal (Le Campion-Alsumard et al., 1993; Mallela and Perry, 2007). Sclerac- tinian corals are the primary carbonate producers on most tropical reefs (Hubbard et al., 1990; Vecsei, 2004), followed by algae, gastropods, bivalves, and foraminiferans (Mallela and Perry, 2007; Perry et al., 2012). Giant clams are rarely mentioned as carbonate contributors to reef frameworks, even though they have large shells, mostly made up of —a calcium carbon- ate polymorph (Moir, 1990). Shell carbonates are generally derived from ambient dissolved inorganic carbon (Romanek and Grossman, 1989), but also include carbon from metabolic respiration and zoo- xanthellae within the mantle tissues (Jones et al., 1986; Watanabe et al., 2004). The relict population of T. gigas from the Great Barrier Reef (Table 2) may only produce 356 kg ha1 yr1 of new shell material but Barker et al. (1988) estimated that a high- density cultured population of T. gigas could produce shell material in excess of 80 t ha1 yr1. The natural T. maxima atoll populations in French Polynesia are capable of producing 23–37 t ha1 yr1 (Table 2) and are so dense that they create small islands called Fig. 2. Fish bite marks on the mantle edge of (a) Tridacna crocea (shell length mapiko (Gilbert et al., 2006). 14 cm) from Tioman Island, Malaysia and (b) Tridacna gigas (shell length 80 cm) from Magnetic Island, Australia. 5.2. Bioeroders 4.3. Hosts for ectoparasites Bioeroders such as grazers, etchers, and borers can increase the Various cyclopoid copepods live within giant clams (Table A2). removal rate of the reef’s carbonate framework (Clapp and Kenk, Even though they are capable of influencing the growth, fecundity, 1963; Hutchings, 1986). The boring giant clam species, T. crocea and survival of their hosts (Finley and Forrester, 2003; Johnson (Hamner and Jones, 1976) and, to a lesser extent, T. maxima et al., 2004), the biology of these cyclopoids is poorly understood. (Yonge, 1980; Hutchings, 1986), are usually found embedded in Anthessius and Lichomolgus are usually found inside the mantle either dead coral heads or dead patches of live colonies (Morton, cavity (Humes, 1972, 1976), while Paclabius inhabits the pericar- 1990). Burrowing by T. crocea has been described as both a dium, i.e. the membrane enclosing the (Kossmann, 1877). mechanical process and chemical etching. Mechanically, T. crocea Multiple cyclopoid species have been found within the same clam enlarge their burrows by grinding back and forth within them, host (Humes, 1972, 1976). Ectoparasitic gastropods are also known and fine shell corrugations on their valves wear away at the bur- to plague giant clams (also see Section 3.2.), and are especially row walls (Yonge, 1953; Hamner and Jones, 1976). Chemical etch- severe in cultured juveniles (Cumming and Alford, 1994). ing (Hedley, 1921; Yonge, 1980) is performed by extending the pedal mantle tissue out of the byssal opening and dissolving the 4.4. Hosts for commensals substrate under and around the clam via excreted solvents (Yonge, 1980; Fatherree, 2006). Given sufficient time and numbers Bivalves host a wide diversity of commensal fauna (Blanco and of settling and growing/burrowing individuals, T. crocea will even- Ablan, 1939; De Grave, 1999), providing refuge (Rosewater, 1965) tually erode away a dead coral head (Hamner and Jones, 1976; and/or food (Fankboner, 1972). The recorded commensals for Glynn, 1997), but this erosive effect is limited to these that tridacnines include pinnotherid pea crabs (Fig. 4; Table A3) and are particular to T. crocea, and does not lead to wide scale attrition pontoniinid shrimps (Fig. 5; Table A4). Pea crabs are common of the reef (Paulay and Kerr, 2001; Aline, 2008). Even though little within the mantle cavities of bivalves (Stauber, 1945; Schmitt is known about the effects of T. maxima’s burrowing, McMichael et al., 1973), positioning themselves on the ctenidial surface () (1974) and Hutchings (1986) both remarked that, due to their with their strong grip and gaining access to food aggregated by the higher densities, they could contribute significantly to biological host (Stauber, 1945). Xanthasia murigera (Fig. 3) is probably the erosion on coral reefs. most widespread, being found in five clam species (Table A3). Pon- toniinid shrimps can also inhabit the mantle cavities of giant 5.3. Topographic enhancement clams. With hooked walking-leg dactyls (Fujino, 1975), they anchor themselves against the currents generated by the gills, Mollusc shells are known to influence their environments, avoiding expulsion (Fankboner, 1972). While some species are either by creating or modifying habitats for other organisms commensal to multiple tridacnine species (Table A4), Anchistus (Gutiérrez et al., 2003). Giant clams can modulate water flow and gravieri appears to be obligate to Hippopus hippopus (McNeill, fluid transport as they add topographical relief to the seabed 1953; Bruce, 1977, 1983) whilst Paranchistus armatus is restricted (Weingarten, 1991; Cabaitan et al., 2008). Depending on their M.L. Neo et al. / Biological Conservation 181 (2015) 111–123 119

Fig. 3. Epibiota diversity amongst giant clam species. (a) Tridacna gigas with a burrowing giant clam (Tridacna crocea) in its shell; Mecherchar Island, Republic of , March 2011. (b) Tridacna derasa with hard coral (Favites sp.) growing on it; Ouvea island of the Loyalty Islands, New Caledonia, August 2010. (c) Hippopus sp. with encrusting crustose coralline algae; Bali, , May 2011. (d) Tridacna maxima hosting a range of encrusting epibionts; Kumejima, Okinawa, Japan; November 2009.

Fig. 4. Commensal pinnotherids (Xanthasia murigera; ZRC2013.0790) found within the mantle cavity of a fluted giant clam (Tridacna squamosa; shell length = 150 mm). (a and b) Carapace length (CL) = 5 mm. (c and d) CL = 11.5 mm.

density, their influence on water flow can be significant. Giant clam 1998). Even after a clam’s death, their heavy valves remain and shells are expected to agitate flow boundary layers much more continue to affect water flow. than the shells of smaller bivalves (Grant et al., 1992; Pilditch et al., 1998), since flow perturbation is correlated to the heights 5.4. Source of zooxanthellae (Symbiodinium spp.) and diameters of protruding objects (Eckman and Nowell, 1984). Aggregations of giant clams are likely to further increase flow Nutrient cycling between zooxanthellae and their coral hosts is perturbation and cause turbulence eddies (Lenihan, 1999). These the key to both organisms’ success in oligotrophic coral reef envi- hydrodynamic disturbances in turn affect the rates at which trans- ronments (Muscatine and Porter, 1977) and similar cycling occurs port of particles and solutes can occur (Gutiérrez et al., 2003). For between zooxanthellae and other organisms, such as zooxanthel- instance, aggregations of bivalves have been shown to alter sedi- late jellyfish (Pitt et al., 2009). Zooxanthellae within giant clams ment transport patterns and rates (Grant et al., 1992; Lenihan, utilize the hosts’ nitrogenous waste with virtually no loss from 1999) and enhance down-flux (Pilditch et al., the system (Hawkins and Klumpp, 1995), meaning that they have 120 M.L. Neo et al. / Biological Conservation 181 (2015) 111–123

6. Conclusion

This review details how giant clams are effective ecosystem engineers that play multiple roles in coral reefs. Their high biomass production, coupled with their wide range of known predators, suggest that giant clams are an important food item. In addition, their gametes and faeces are food to opportunistic feeders. Due to their large shell size, giant clams can shelter reef fish as well as support a diverse and abundant array of epibionts, ectoparasites, and commensals. Furthermore, some species, such as the pontonii- nid (Paranchistus armatus), are only found in tridacnine hosts. At the reef-scale, dense populations of giant clams can annu- ally contribute 100s to 1,000s kg ha1 of shell material to a coral reef, far outweighing localized erosion by T. crocea and T. maxima. Giant clams provide their symbiotic Symbiodinium with nutrients and protection, resulting in tridacnines acting as algal ‘reservoirs’. They also filter large volumes of water—which can potentially Fig. 5. A commensal pontoniinid (Anchistus sp.; body length = 34 mm) found counteract eutrophication. While we have only evaluated the eco- resting on the mantle of a fluted giant clam (Tridacna squamosa; shell length = 243 mm). logical roles of the more common giant clam species: H. hippopus, T. crocea, T. derasa, T. gigas, T. maxima, and T. squamosa, we expect that the rarer species perform functions similar to those of their far greater access to nitrogen than they would if living in the close relatives. surrounding seawater. Giant clams also protect their symbiotic Even though there have been numerous giant clam population zooxanthellae from predation (Fankboner, 1971) and excessive collapses (e.g. Munro, 1989; Kinch and Teitelbaum, 2010; Neo ultraviolet irradiation (Ishikura et al., 1997). and Todd, 2012), it is difficult to measure the ecosystem-level Giant clams release large numbers of zooxanthellae in their fae- effects of these events due to the concomitant impacts of multiple cal pellets. To regulate symbiont density a T. derasa can discharge stressors that typify contemporary reefs (Hughes and Connell, 4.9 105 cells clam1 d1 of intact zooxanthellae (Maruyama 1999). Nevertheless, some negative consequences are predictable, and Heslinga, 1997) and T. gigas can discharge 4.7 105 - for example, biomass and carbonate production, surface area for cells clam1 d1 (Buck et al., 2002). These are both several orders epibionts, and water filtering, are all expected to decrease with of magnitude higher than the release rates of corals (Yamashita reduced giant clam abundance. Other effects might require thresh- et al., 2011). As noted by Maruyama and Heslinga (1997, p.475), olds to be breached, for example, a minimum density of giant ‘‘most of the discharged zooxanthellae were indistinguishable from clams may be needed before they act as effective fish nurseries. intact algal cells freshly isolated from the mantle.’’ Trench et al. It is unlikely, however, that researchers would experimentally (1981) also found that zooxanthellae in faecal pellets were intact, remove clams from a healthy reef to measure the outcome. On photosynthetically active, and culturable. The branched tubular the other hand, restocking programmes present an excellent system extending from a giant clam’s into its mantle opportunity to monitor the response of the reef to enhanced clam (Fankboner and Reid, 1990; Norton et al., 1992) provides numerous numbers; unfortunately this is rarely done. As mentioned in the microhabitats for zooxanthellae (Norton et al., 1992) allowing mul- Introduction, the one exception is Cabaitan et al. (2008), who tiple clades or multiple types from a single clade of symbionts to specifically set out to test the localized effects of transplanting co-exist in a single host (Baillie et al., 2000; DeBoer et al., 2012). maricultured giant clams (T. gigas, >40 cm shell length) into repli- The substantial quantities and (possibly) types of zooxanthellae cate 5 5m2 plots of degraded patch reef in the Philippines (25 released from giant clams become available for other zooxanthel- clams per plot). Within three months, their ‘‘other biota’’ category late-dependent species to ‘take up’, hence contributing to the (i.e. ascidians, anemones, gorgonians, soft corals, sponges, and wider coral reef ecosystem. zoanthids) had increased significantly from 2.0% to 14.8% cover; with no change observed in the control plots. Within the same time period fish species richness and abundance also increased sig- 5.5. Counteractors of eutrophication nificantly. Similar research at different locations, and with other giant clam species and densities, is needed. In coastal marine waters, corals may be competitively excluded Any significant ecological benefits will likely only accrue where by macroalgae or heterotrophic filter feeders as the water becomes giant clams are present in healthy, i.e. self-sustaining, populations more eutrophic (Fabricius, 2005). Shallow water benthic bivalves and hence their conservation is paramount. As highlighted by Neo are known to be natural controllers of eutrophication (Officer and Todd (2013), the CITES and IUCN data for giant clams are out- et al., 1982) and giant clams can perform this function in two ways: dated and potentially misleading. Importantly, there are now three by filtering water and by sequestering nutrients (Klumpp and species recently rediscovered or undescribed, plus one entirely Griffiths, 1994). Giant clams filter large quantities of seawater; new species, that have no official conservation status (Table 1). even a sparse population of mature T. gigas (0.04 clams m2)on Giant clams should feature more prominently when planning mar- the Great Barrier Reef is capable of filtering over 28,000 l ha1 h1 ine protected areas and integrated schemes (Table 2). Giant clams also clear water of algal cells efficiently, e.g. (van Wynsberge et al., 2013) and national/local assessments must Tridacna species ingest 51–58% while H. hippopus ingests 81% be part of this process. Notably, not only are giant clams useful for (Klumpp and Griffiths, 1994). Whether algal biomass is assimilated the functioning of coral reefs, they can help protect them by acting by the clams or excreted as faeces, it is removed from the water as surrogate species. Giant clams are already considered an indica- column in the short term and will therefore not contribute to tur- tor species by Reef Check (Hodgson, 2001) and, being well-known bidity. By locking assimilated nutrients away in their biomass charismatic invertebrate megafauna (for instance, fourteen cul- (Table 2), giant clams sequester them from the water where they tured T. gigas are used in snorkel trails at Magnetic Island; Braley, could otherwise encourage macroalgae to flourish. pers. comm., 2014), they have the potential to play a flagship role M.L. Neo et al. / Biological Conservation 181 (2015) 111–123 121

(sensu Walpole and Leader-Williams, 2002, but see Favreau et al., Ault, T.R., Johnson, C.R., 1998. Relationships between habitat and recruitment of 2006) in reef conservation. three species of damselfish (Pomacentridae) at Heron Reef, Great Barrier Reef. J. Exp. Mar. Biol. Ecol. 223, 145–166. We are not proposing that giant clams are essential to the Baillie, B.K., Belda-Baillie, C.A., Maruyama, T., 2000. Conspecificity and Indo-Pacific survival of coral reefs; however, there can be no doubt that they distribution of Symbiodinium genotypes (Dinophyceae) from giant clams. J. make a positive contribution to these critically important tropical Phycol. 36, 1153–1161. Barker, J.R., Crawford, C.M., Shelley, C., Braley, R.D., Lucas, J.S., Nash, W.J., Lindsay, S., ecosystems. Based on the wide range of ecological functions they 1988. Ocean-nursery technology and production data from lines and cover for perform, giant clams are unique among reef organisms and there- the giant clam, Tridacna gigas. In: Copland, J.W., Lucas, J.S. (Eds.), Giant clams in fore deserve attention. In combination with their status as the Asia and the Pacific, Australian Centre for International Agricultural Research, Canberra, Australia. Monograph No. 9, pp. 225–228. world’s largest bivalves and their popularity with SCUBA divers, a Bell, J.D., Galzin, R., 1984. Influence of live coral cover on a coral reef fish greater understanding of giant clams’ contributions will provide communities. Mar. Ecol. Prog. Ser. 15, 265–274. managers with ‘ammunition’ to justify their protection. Crucially, Beukers, J.S., Jones, G.P., 1997. Habitat complexity modifies the impact of piscivores on a coral reef fish population. Oecologia 114, 50–59. both their international and local conservation statuses need to bin Othman, A.S., Goh, G.H.S., Todd, P.A., 2010. The distribution and status of giant be updated and monitored (Brito et al., 2010; Neo and Todd, clams (family Tridacnidae) – a short review. Raffles Bull. Zool. 58, 103–111. 2013) if appropriate management strategies are to be developed. Black, R., Johnson, M.S., Prince, J., Brearley, A., Bond, T., 2011. Evidence of large, local Whatever safeguards can be established will not only boost giant variations in recruitment and mortality in the small giant clam, Tridacna maxima, at Ningaloo Marine Park, Western Australia. Mar. Freshwater Res. 62, clam populations but, by extension, also benefit coral reefs. 1318–1326. Blanco, G.J., Ablan, G.L., 1939. A rare parasitic new to Pangasinan Province, Acknowledgements Luzon. Philipp. J. Sci. 70, 217–219. Boglio, E.C., Lucas, J.S., 1997. Impacts of ectoparasitic gastropods on growth, survival, and physiology of juvenile giant clams (Tridacna gigas), including a We would like to thank those who kindly contributed simulation model of mortality and reduced growth rate. 150, 25– photographs: Dr. Rick Braley (Fig. 2b), Tan Hong Yee (Fig. 3a), Dr. 43. Borsa, P., Fauvelot, C., Tiavouane, J., Grulois, D., Wabnitz, C., Abdon Naguit, M.R., Tan Heok Hui (Fig. 3b and d) and Jeffrey Low (Fig. 3c). Ali Eimran, Andréfouët, S., 2014. Distribution of Noah’s giant clam, Tridacna noae. Mar. Serina Lee, Lee Yen-Ling, Dr. Laurence Liao, Lim Swee Cheng and . http://dx.doi.org/10.1007/s12526-014-0265-9. Dr. Tan Swee Hee assisted with epibiont identification, and Profes- Botton, M.L., 2009. The ecological importance of horseshoe crabs in estuarine and coastal communities: a review and speculative summary. In: Tanacredi, J.T., sor Peter Ng Kee Lin and Dr. Sammy De Grave assisted with iden- Botton, M.L., Smith, D. (Eds.), Biology and Conservation of Horseshoe Crabs. tification of pea crabs and shrimps respectively. Two anonymous Springer-Verlag, US, pp. 45–63. reviewers and Dr. Rick Braley reviewed this paper and provided Braley, R.D., 1984. Reproduction in the giant clams Tridacna gigas and T. derasa in situ on the North-Central Great Barrier Reef, Australia, and Papua New input that greatly improved the final manuscript. This research Guinea. Coral Reefs 3, 221–227. was supported by National Parks Board CME Grant number R- Braley, R.D., 1987. Distribution and abundance of the giant clams Tridacna gigas and 154-000-568-490. Dr. Neo Mei Lin also acknowledges the World T. derasa on the Great Barrier Reef. Micronesica 20, 215–223. Future Foundation for her PhD Prize in Environmental and Sustain- Bridge, T.C.L., Hughes, T.P., Guinotte, J.M., Bongaerts, P., 2013. Call to protect all coral reefs. Nat. Clim. Change 3, 528–530. ability Research 2013. Brito, D., Ambal, R.G., Brooks, T., De Silva, N., Foster, M., Wang, H., Hilton-Taylor, C., Paglia, A., Rodríguez, J.P., Rodríguez, J.V., 2010. How similar are national red lists and the IUCN red list? Biol. Conserv. 143, 1154–1158. Appendix A. Supplementary material Bruce, A.J., 1977. Pontoniine shrimps in the collections of the Australian museum. Rec. Aust. Mus. 31, 39–81. Supplementary data associated with this article can be found, in Bruce, A.J., 1983. The pontoniine shrimp fauna of Australia. In: Lowry, J.K. (Ed.), Papers from the Conference on the Biology and Evolution of Crustacea. The the online version, at http://dx.doi.org/10.1016/j.biocon.2014.11. Australian Museum, Sydney, New South Wales, Australia, pp. 195–218. 004. Bruce, A.J., 2000. Biological observations on the commensal shrimp Paranchistus armatus (H. Milne Edwards) (Crustacea: : Pontoniinae). Beagle (Rec. Mus. Art Galleries Northern Terr.) 16, 91–96. Buck, B.H., Rosenthal, H., Saint-Paul, U., 2002. Effect of increased irradiance and References thermal stress on the symbiosis of Symbionidium microadriaticum and Tridacna gigas. Aquat. Living Resour. 15, 107–117. Bustard, R., 1972. Sea Turtles, their Natural History and Conservation. Collins, Abdon-Naguit, M.R., Alcazar, S.N., 1989. Notes on predation on giant clams London. (: Tridacnidae). In: Zaragoza, E.C., de Guzman, D.L., Gonzales, E.P. Cabaitan, P.C., Gomez, E.D., Aliño, P.M., 2008. Effects of coral transplantation and (Eds.), Culture of giant clams (Bivalvia: Tridacnidae). Proceedings of the giant clam restocking on the structure of fish communities on degraded patch Symposium on the Culture of Giant Clams, Silliman University, Dumaguete reefs. J. Exp. Mar. Biol. Ecol. 357, 85–98. City, March 15–17 1988. Philippine Council for Aquatic and Marine Research Carter, R.M., 1968. On the biology and palaeontology of some predators of bivalved and Development and Australian Centre for International Agricultural Research . Palaeogeogr. Palaeoclimatol. 4, 29–65. (Book Series No. 02/1989), pp. 81–83. Chambers, C.N.L., 2007. Pasua (Tridacna maxima) size and abundance in Tongereva Abellö, P., Villanueva, R., Gili, J.M., 1990. Epibiosis in deep-sea crab populations as , Cook Islands. SPC Trochus Inform. Bull. 13, 7–12. indicators of biological and behavioural characteristics of the host. J. Mar. Biol. Chan, S.W., 2007. Ontogenetic changes in feeding ecology and habitat of the Assoc. UK 70, 687–695. damselfish Neoglyphidodon melas at Lizard Island, Great Barrier Reef. Accordi, G., Brilli, M., Carbone, F., Voltaggio, M., 2010. The raised coral reef complex Independent Study Project (ISP) Collection, Paper 146. of the Kenyan coast: Tridacna gigas U-series dates and geological implications. J. Chantrapornsyl, S., Kittiwattanawong, K., Adulyanukosol, K., 1996. Distribution and Afr. Earth Sci. 58, 97–114. abundance of giant clam around Lee-Pae Island, the Andaman Sea, Thailand. Adams, A.L., Needham, E.W., Knauer, J., 2013. The effect of shade on water quality Phuket Mar. Biol. Center Spec. Publ. 16, 195–200. parameters and survival and growth of juvenile fluted giant clams, Tridacna Chisholm, J.R.M., 2003. Primary productivity of reef-building crustose coralline squamosa, cultured in a land-based growth trial. Aquacult. Int. 21, 1311–1324. algae. Limnol. Oceanogr. 48, 1376–1387. Alcazar, S.N., 1986. Observations on predators of giant clams (Bivalvia: family Clapp, W.F., Kenk, R., 1963. Marine Borers. An Annotated Bibliography. Office of Tridacnidae). Silliman J. 33, 54–57. Navy Research, Department of the Navy, Washington, DC. Aline, T., 2008. Dissolution of dead corals by euendolithic microorganisms across Creed, J.C., 2000. Epibiosis on cerith shells in a seagrass bed: correlation of shell the northern Great Barrier Reef (Australia). Microb. Ecol. 55, 569–580. occupant with epizoite distribution and abundance. Mar. Biol. 137, 775– Andréfouët, S., Gilbert, A., Yan, L., Remoissenet, G., Payri, C., Chancerelle, Y., 2005. 782. The remarkable population size of the endangered clam Tridacna maxima Cumming, R.L., Alford, R.A., 1994. Population dynamics of Turbonilla sp. assessed in Fangatau Atoll (Eastern Tuamotu, French Polynesia) using in situ (Pyramidellidae, Opisthobranchia), an ectoparasite of giant clams in and remote sensing data. ICES J. Mar. Sci. 62, 1037–1048. mariculture. J. Exp. Mar. Biol. Ecol. 183, 91–111. Andréfouët, S., Van Wynsberge, S., Gaertner-Mazouni, N., Menkes, C., Gilbert, A., De Grave, S., 1999. Pontoniinae (Crustacea: Decapoda: Palaemonidae) associated Remoissenet, G., 2013. Climate variability and massive mortalities challenge with bivalve molluscs from Hansa Bay, Papua New Guinea. Bull. Inst. Roy. Sci. giant clam conservation and management efforts in French Polynesia atolls. Nat. Belg.-Biol. 69, 125–141. Biol. Conserv. 160, 190–199. DeBoer, T.S., Baker, A.C., Erdmann, M.V., Ambariyanto, Jones, P.R., Barber, P.H., 2012. Arvedlund, M., Takemura, A., 2005. Long-term observation in situ of the Patterns of Symbionidium distribution in three giant clam species across the anemonefish Amphiprion clarkii (Bennett) in association with a soft coral. biodiverse Bird’s Head region of Indonesia. Mar. Ecol. Prog. Ser. 444, 117–132. Coral Reefs 24, 698. 122 M.L. Neo et al. / Biological Conservation 181 (2015) 111–123

Delbeek, J.C., Sprung, J., 1994. The Reef , vol. 1. Ricordea Publishing, Hodgson, G., 2001. Reef check: the first step in community-based management. Coconut Grove, FL. Bull. Mar. Sci. 69, 861–868. Dumas, P., Tiavouane, J., Senia, J., William, A., Dick, L., Fauvelot, C., 2014. Evidence of Huang, D., Todd, P.A., Guest, J.R., 2007. Movement and aggregation in the fluted early chemotaxis contributing to active habitat selection by the sessile giant giant clam (Tridacna squamosa L.). J. Exp. Mar. Biol. Ecol. 342, 269–281. clam Tridacna maxima. J. Exp. Mar. Biol. Ecol. 452, 63–69. Hubbard, D.K., Miller, A.I., Scaturo, D., 1990. Production and cycling of calcium Eckman, J.E., Nowell, A.R.M., 1984. Boundary skin friction and sediment transport carbonate in a shelf-edge reef system (St. Croix, U.S. Virgin Islands): about an -tube mimic. Sedimentology 31, 851–862. applications to the nature of reef systems in the fossil record. J. Sediment. Fabricius, K.E., 2005. Effects of terrestrial runoff on the ecology of corals and coral Petrol. 60, 335–360. reefs: review and synthesis. Mar. Pollut. Bull. 50, 125–146. Huelsken, T., Keyse, J., Liggins, L., Penny, S., Treml, E.A., Riginos, C., 2013. A novel Fankboner, P.V., 1971. Intracellular digestion of symbiotic zooxanthellae by host widespread cryptic species and phylogeographic patterns within several giant amoebocytes in giant clams (Bivalvia: Tridacnidae), with a note on the nutritional clam species (Cardiidae: Tridacna) from the Indo-Pacific Ocean. PLoS ONE 8, role of the hypertrophied siphonal epidermis. Biol. Bull. 141, 222–234. e80858. http://dx.doi.org/10.1371/journal.pone.0080858. Fankboner, P.V., 1972. On the association between the pontonid shrimp Anchistus Hughes, T.P., Connell, J.H., 1999. Multiple stressors on coral reefs: a long-term miersi de Man (Decapoda, Palaemonidae) and giant clams (Lamellibranchia, perspective. Limnol. Oceanogr. 44, 932–940. Tridacnidae). Wasmann J. Biol. 30, 35–42. Humes, A.G., 1972. Cyclopoid copepods associated with Tridacnidae (Mollusca, Fankboner, P.V., Reid, R.G.B., 1990. Nutrition in giant clams (Tridacnidae). In: Bivalvia) at Eniwetok Atoll. Proc. Biol. Soc. Wash. 84, 345–358. Morton, B. (Ed.), The Bivalvia–Proceedings of a Memorial Symposium in Honour Humes, A.G., 1976. Cyclopoid copepods associated with Tridacnidae (Mollusca, of Sir Charles Maurice Yonge, Edinburgh, 1986. Hong Kong University Press, Bivalvia) in the Moluccas. Proc. Biol. Soc. Wash. 89, 491–508. Hong Kong, pp. 195–209. Hutchings, P.A., 1986. Biological destruction of coral reefs. Coral Reefs 4, 239–252. Fatherree, J.W., 2006. Giant Clams in the Sea and the Aquarium: The biology, Hviding, E. (Ed.), 1993. The Rural Context of Giant Clam Mariculture in the Solomon Identification, and Aquarium Husbandry of Tridacnid clams. Liquid Medium Islands: An Anthropological Study. International Center for Living Aquatic Publications. Resources Management, Manila, Philippines. Favreau, J.M., Drew, C.A., Hess, G.R., Rubino, M.J., Koch, F.H., Eschelbach, K.A., 2006. Ishikura, M., Kato, C., Maruyama, T., 1997. UV-absorbing substances in Recommendations for assessing the effectiveness of surrogate species zooxanthellate and azooxanthellate clams. Mar. Biol. 128, 649–655. approaches. Biodivers. Conserv. 15, 3949–3969. Jantzen, C., Wild, C., El-Zibdah, M., Roa-Quiaoit, H., Haacke, C., Richter, C., 2008. Feiferak, B.P., 1987. Spines and epibionts as antipredator defenses in the thorny Photosynthetic performance of giant clams, Tridacna maxima and T. squamosa, americanus Hermann. J. Exp. Mar. Biol. Ecol. 105, 39–56. Red Sea. Mar. Biol. 155, 211–221. Finley, R.J., Forrester, G.E., 2003. Impact of ectoparasites on the demography of a Johnson, S.C., Treasurer, J.W., Bravo, S., Nagasawa, K., Kabata, Z., 2004. A review of small reef fish. Mar. Prog. Ecol. Ser. 248, 305–309. the impact of parasitic copepods on marine aquaculture. Zool. Stud. 43, 229– Fujino, T., 1975. Fine features of the dactylus of the ambulatory pereiopods in a 243. bivalve-associated shrimp, Anchistus miersi (De Man), under the scanning Jones, D.S., Williams, D.F., Romanek, C.S., 1986. Life history of symbiont-bearing electron microscope (Decapoda, Natantia, Pontoniinae). Crustaceana 29, 252– giant clams from stable isotope profiles. Science 231, 46–48. 254. Keable, S.J., 1995. Structure of the marine invertebrate scavenging guild of a tropical Gilbert, A., Andréfouët, S., Yan, L., Remoissenet, G., 2006. The giant clam Tridacna reef ecosystem – field studies at Lizard Island, Queensland, Australia. J. Nat. maxima communities of three French Polynesia islands: comparison of their Hist. 29, 27–45. population sizes and structures at early stages of their exploitation. ICES J. Mar. Kinch, J., Teitelbaum, A., 2010. Proceedings of the regional workshop on the Sci. 63, 1573–1589. management of sustainable fisheries for giant clams (Tridacnidae) and CITES Glynn, P.W., 1997. and coral reef growth: a dynamic balance. In: capacity building. 4–7 August 2009, Nadi, Fiji. Secretariat of the Pacific Birkeland, C. (Ed.), Life and Death of Coral Reefs. Chapman and Hall, New York, Community, Coastal Fisheries Programme, Noumea, New Caledonia. pp. 68–95. Klumpp, D.W., Griffiths, C.L., 1994. Contributions of phototrophic and heterotrophic Govan, H., 1992a. Predators of maricultured Tridacnid clams. In: Richmond, R.H. nutrition to the metabolic and growth requirements of four species of giant (Ed.), Proceedings of the 7th International Coral Reef Symposium, Vol. 2. clam (Tridacnidae). Mar. Prog. Ecol. Ser. 115, 103–115. University of Press, UOG Station, Guam, pp. 739–743. Klumpp, D.W., Bayne, B.L., Hawkins, A.J.S., 1992. Nutrition of the giant clam Tridacna Govan, H., 1992b. Predators and predator control. In: Calumpong, H.P. (Ed.), The gigas (L.). I. Contribution of filter-feeding and photosynthates to respiration and Giant Clam: An Ocean Culture Manual. Australian Centre for International growth. J. Exp. Mar. Biol. Ecol. 155, 105–122. Agricultural Research, Monograph No. 16, Canberra, Australia, pp. 41–49. Kossmann, R., 1877. Entomostraca. (1. Theil: Licholmolgidae). Zool Ergebnisse einer Govan, H., 1995. Cymatium muricinum and other ranellid gastropods: major im Auftrage de Königl Akad Wiss Berlin ausgeführten Reise in die Küstengebiete predators of cultured Tridacnid clams. ICLARM Technical Reports 49. des Rothen Meeres, erste Hälfte 4, pp. 1–24 (in German). Govan, H., Fabro, L.Y., Ropeti, E., 1993. Controlling predators of cultured Tridacnid Le Campion-Alsumard, T., Romano, J.-C., Peyrot-Clausade, M., Le Campion, J., Paul, clams. In: Fitt, W.K. (Ed.), Biology and Mariculture of Giant Clams. A workshop R., 1993. Influence of some coral reef communities on the calcium carbonate held in conjunction with the 7th International Coral Reef Symposium, 21–26 budget of Tiahura reef (Moorea, French Polynesia). Mar. Biol. 115, 685–693. June 1992, Guam, USA, pp. 111–118. Lecchini, D., Planes, S., Galzin, R., 2007. The influence of habitat characteristics and Grant, J., Emerson, C.W., Shumway, S.E., 1992. Orientation, passive transport, and conspecifics on attraction and survival of coral reef fish juveniles. J. Exp. Mar. sediment erosion features of the sea magellanicus in the Biol. Ecol. 341, 85–90. . Can. J. Zool. 71, 953–959. Lenihan, H.S., 1999. Physical-biological coupling on oyster reefs: how habitat Gratwicke, B., Speight, M.R., 2005. The relationship between fish species richness, structure influences individual performance. Ecol. Monogr. 69, 251–275. abundance and habitat complexity in a range of shallow tropical marine Lescureux, N., Linnell, J.D.C., 2014. Warring brothers: the complex interactions habitats. J. Fish Biol. 66, 650–667. between wolves (Canis lupus) and dogs (Canis familiaris) in a conservation Guest, J.R., Todd, P.A., Goh, E., Sivalonganathan, B.S., Reddy, K.P., 2008. Can giant context. Biol. Conserv. 171, 232–245. clams (Tridacna squamosa) populations be restored in Singapore’s heavily Ling, H., 2007. The defensive role of scutes in juvenile fluted giant clams (Tridacna impacted coral reefs? Aquat. Conserv. 18, 570–579. squamosa). Unpublished Honours Thesis, Department of Biological Sciences, Gutiérrez, J.L., Jones, C.G., Strayer, D.L., Iribarne, O.O., 2003. Mollusks as ecosystem National University of Singapore. engineers: the role of shell production in aquatic habitats. Oikos 101, 79–90. Ling, H., Todd, P.A., Chou, L.M., Yap, V.B., Sivalonganathan, B., 2008. The defensive Hamner, W.M., Jones, M.S., 1976. Distribution, burrowing, and growth rates of the role of scutes in juvenile fluted giant clams (Tridacna squamosa). J. Exp. Mar. clam Tridacna crocea on interior reef flats. Oecologia 24, 207–227. Biol. Ecol. 359, 77–83. Harder, T., 2008. Marine epibiosis: concepts, ecological consequences and host Lingo, M.E., Szedlmayer, S.T., 2006. The influence of habitat complexity on reef fish defence (Chapter 12). In: Flemming, H.-C., Murthy, P.S., Venkatesan, R., Cooksey, communities in the northeastern Gulf of Mexico. Environ. Biol. Fish. 76, 71–80. K.E. (Eds.), Marine and Industrial Biofouling (Springer Series on Biofilms), vol. 4, Loch, I., 1991. Melo meals. Aust. Shell News 73, 5. first ed. Springer-Verlag Berlin Heidelberg, pp. 219–232. Lucas, J.S., 1994. The biology, exploitation, and mariculture of giant clams Hardy, J.T., Hardy, S.A., 1969. Ecology of Tridacna in Palau. Pac. Sci. 23, 467–472. (Tridacnidae). Rev. Fish. Sci. 2, 181–223. Harzhauser, M., Mandic, O., Piller, W.E., Reuter, M., Kroh, A., 2008. Tracing back the Maboloc, E.A., Mingoa-Licuanan, S.S., 2011. Feeding aggregation of Spratelloides origin of the Indo-Pacific mollusc fauna: basal Tridacninae from the delicatulus on giant clams’ gametes. Coral Reefs 30, 167. and Miocene of the Sultanate of Oman. Palaeontology 51, 199–213. Mallela, J., Perry, C.T., 2007. Calcium carbonate budgets for two coral reefs affected by Hawkins, A.J.S., Klumpp, D.W., 1995. Nutrition of the giant clam Tridacna gigas (L.) II. different terrestrial runoff regimes, Rio Bueno, Jamaica. Coral Reefs 26, 129–145. Relative contributions of filter-feeding and the ammonium-nitrogen acquired Maruyama, T., Heslinga, G.A., 1997. Fecal discharge of zooxanthellae in the giant and recycled by symbiotic alga towards total nitrogen requirements for tissue- clam Tridacna derasa, with reference to their in situ growth rate. Mar. Biol. 127, growth and . J. Exp. Mar. Biol. Ecol. 190, 263–290. 473–477. Hedley, C., 1921. A revision of the Australian Tridacna. Rec. Aust. Mus. 13, 163–172. McMichael, D.F., 1974. Growth rate, population size and mantle coloration in the Heslinga, G.A., 2013. Saving giants (eBook). Cultivation and conservation of small giant clam Tridacna maxima (Röding), at One Tree Island, Capricorn Group, Tridacnid clams. . Queensland. In: Cameron, A.M., Campbell, B.M., Cribb, A.B., Endean, R., Jell, J.S., Heslinga, G.A., Fitt, W.K., 1987. The domestication of reef-dwelling clams. Jones, O.A., Mather, P., Talbot, F.H. (Eds.), Proceedings of the 2nd International BioScience 37, 332–339. Coral Reef Symposium, 1. The Great Barrier Reef Committee, Brisbane, Australia, Heslinga, G.A., Perron, F.E., Orak, O., 1984. Mass culture of giant clams (F. pp. 241–254. Tridacnidae) in Palau. Aquaculture 39, 197–215. McNeill, F.A., 1953. Carcinological notes No. 2. Rec. Aust. Mus. 23, 89–96. Heslinga, G.A., Watson, T.C., Isamu, T., 1990. Giant Clam Farming. Pacific Fisheries Mercier, A., Hamel, J.-F., 1996. The secret of the giant clam. Freshwater Mar. Development Foundation (NMFS/NOAA), , Hawaii, USA. Aquarium 19, 112–113. M.L. Neo et al. / Biological Conservation 181 (2015) 111–123 123

Miller, D., 1979. National sites survey: a summary report. Solomon Islands National and Mermaid Reef, northwestern Australia. Report to Environment Australia, Museum, Honiara Australian Institute of Marine Science, Townsville, Queensland. Mingoa-Licuanan, S.S., Gomez, E.D., 2002. Giant clam conservation in Southeast Ricard, M., Salvat, B., 1977. Faeces of Tridacna maxima (Mollusca: Bivalvia), Asia. Trop. Coasts 3, 24–56. composition and coral reef importance. In: Taylor, D.L. (Ed.), Proceedings of Moir, B.G., 1990. Comparative studies of ‘‘fresh’’ and ‘‘aged’’ Tridacna gigas shell: the 3rd International Coral Reef Symposium, vol. 1. Biology. Rosenstiel School of Preliminary investigation of a reported technique for pretreatment of tool Marine and Atmospheric Science, Miami, Florida, pp. 495–501. material. J. Archaeol. Sci. 17, 329–345. Ricciardi, A., Bourget, E., 1998. Weight-to-weight conversion factors for marine Molluscs Specialist Group, 1996. Tridacna crocea. In: IUCN 2013. IUCN Red List of benthic macroinvertebrates. Mar. Ecol. Prog. Ser. 163, 245–251. Threatened Species. Version 2013.2. (accessed 05.05.14). Takapoto Lagoon (French Polynesia). In: Taylor, D.L. (Ed.), Proceedings of the 3rd Morton, B., 1990. Corals and their bivalve borers—the evolution of a symbiosis. In: International Coral Reef Symposium, vol. 1. Biology. Rosenstiel School of Marine Morton, B. (Ed.), The Bivalvia—Proceedings of a Memorial Symposium in and Atmospheric Science, Miami, Florida, pp. 599–605. Honour of Sir Charles Maurice Yonge, Edinburgh, 1986. Hong Kong University Richardson, W.J., 1991. Predation of juvenile tridacnid clams in two high-island Press, Hong Kong, pp. 11–46. fringing reefs in the Great Barrier Reef. Unpublished Honours thesis, Munro, J.L., 1989. Fisheries for giant clams (Tridacnidae: Bivalvia) and prospects for Department of , James Cook University of Northern Queensland, stock enhancement. In: Caddy, J.F. (Ed.), Marine Invertebrate Fisheries: Their Australia. Assessment and Management. John Wiley and Sons Inc., New York, pp. 541– Richter, C., Roa-Quiaoit, H., Jantzen, C., Al-Zibdah, M., Kochzius, M., 2008. Collapse of 558. a new living species of giant clam in the Red Sea. Curr. Biol. 18, 1349–1354. Muscatine, L., Porter, J.W., 1977. Reef Corals: Mutualistic Symbioses Adapted to Rogers, C.S., Salesky, N.H., 1981. Productivity of Acropora palmata (Lamarck), Nutrient-Poor Environments. Bioscience 27, 454–460. macroscopic algae, and algal turf from Tague Bay Reef, St. Croix, U.S. Virgin Naumann, M.S., Jantzen, C., Haas, A.F., Iglesias-Prieto, R., Wild, C., 2013. Benthic Islands. J. Exp. Mar. Biol. Ecol. 49, 179–187. primary production budget of a Caribbean reef lagoon (Puerto Morelos, Romanek, C.S., Grossman, E.L., 1989. Stable isotope profiles of Tridacna maxima as Mexico). PLoS ONE 8 (12), e82923. http://dx.doi.org/10.1371/journal.pone. environmental indicators. Palaios 4, 402–413. 0082923. Roscoe, E.H., 1962. Some records of large Tridacna specimens. Hawaiian Shell News Neo, M.L., Todd, P.A., 2011a. Predator-induced changes in fluted giant clam 11, 8. (Tridacna squamosa) shell morphology. J. Exp. Mar. Biol. Ecol. 397, 21–26. Rosewater, J., 1965. The family Tridacnidae in the Indo-Pacific. Indo-Pac. Mollusca 1, Neo, M.L., Todd, P.A., 2011b. Quantification of water squirting by juvenile fluted 347–396. giant clams (Tridacna squamosa L.). J. Ethol. 29, 85–91. Schmitt, W.L., McCain, J.C., Davidson, E.S., 1973. Crustaceorum Catalogus Pars 3. Neo, M.L., Todd, P.A., 2012. Giant clams (Mollusca: Bivalvia: Tridacninae) in Decapoda I, Brachyura I, Fam. . Dr. W. Junk, Den Haag. Singapore: history, research and conservation. Raffles B. Zool. 25, 67–78. Soo, P., Todd, P.A., 2014. The behaviour of giant clams (Bivalvia: Cardiidae: Neo, M.L., Todd, P.A., 2013. Conservation status reassessment of giant clams Tridacninae). Mar. Biol. 161, 2699–2717. (Mollusca: Bivalvia: Tridacninae) in Singapore. Nat. Singapore 6, 125–133. Stauber, L.A., 1945. Pinnotheres ostreum, parasitic on the American oyster Neo, M.L., Erftemeijer, P.L.A., van Beek, J.K.L., van Maren, D.S., Teo, S.L.-M., Todd, P.A., () virginica. Biol. Bull. 88, 269–291. 2013. Recruitment constraints in Singapore’s fluted giant clam (Tridacna Su, Y., Hung, J.-H., Kubo, H., Liu, L.-L., 2014. Tridacna noae (Röding, 1798) – a valid squamosa) population – a dispersal model approach. PLoS ONE 8 (3), e58819. giant clam species separated from T. maxima (Röding, 1798) by morphological http://dx.doi.org/10.1371/journal.pone.0058819. and genetic data. Raffles Bull. Zool. 62, 124–135. Newman, L.J., Cannon, L.R.G., Govan, H., 1991. Clam killers – Turbellarian predators Todd, P.A., Lee, J.H., Chou, L.M., 2009. Polymorphism and crypsis in the boring giant or scavengers? Australian Marine Science Association Conference, July 8–12, clam (Tridacna crocea): potential strategies against visual predators. 1991, Brisbane, Australia (Abstract). Hydrobiologia 635, 37–43. Newman, L.J., Cannon, L.R.G., Govan, H., 1993. Stylochus (Imogene) matatasi n. sp. Trench, R.K., Wethey, D.S., Porter, J.W., 1981. Observations on the symbiosis with (Platyhelminthes, Polycladida): pest of cultured giant clams and zooxanthellae among the Tridacnidae (Mollusca, Bivalvia). Biol. Bull. 161, 180– from Solomon Islands. Hydrobiologia 257, 185–189. 198. Norton, J.H., Shepherd, M.A., Long, H.M., Fitt, W.K., 1992. The zooxanthellal tubular Trott, L.B., Chan, W.L., 1972. Carapus homei commensal in mantle cavity of Tridacna system in the giant clam. Biol. Bull. 183, 503–506. sp. in South Sea. Copeia 4, 872–873. Norton, J.H., Shepherd, M.A., Long, H.M., Prior, H.C., 1993. Parasites of the giant Turner, R.D., Boss, K.J., 1962. The in the Western Atlantic. clams (Tridacnidae). In: Fitt, W.K. (Ed.) Biology and Mariculture of Giant Clams. Johnsonia 4, 81–116. Australian Centre for International Agricultural Research, Proceedings No. 47, van Wynsberge, S., Andréfouët, S., Gilbert, A., Stein, A., Remoissenet, G., 2013. Best Canberra, Australia, pp. 18–23. management strategies for sustainable giant clam fishery in French Polynesia Officer, C.B., Smayda, T.J., Mann, R., 1982. Benthic filter feeding: a natural Islands: answers from a spatial modeling approach. PLoS ONE 8, e64641. http:// eutrophication control. Mar. Ecol. Prog. Ser. 9, 203–210. dx.doi.org/10.1371/journal.pone.0064641. Ono, R., Clark, G., 2012. A 2500-year record of marine resource use on Ulong Island, Vecsei, A., 2004. A new estimate of global reefal carbonate production including the Republic of Palau. Int. J. Osteoarchaeol. 22, 637–654. fore-reefs. Global Planet. Change 43, 1–18. Paulay, G., Kerr, A., 2001. Patterns of coral reef development in Tarawa Atoll Vicentuan-Cabaitan, K., Neo, M.L., Eckman, W., Teo, S.L.-M., Todd, P.A., 2014. Giant (). Bull. Mar. Sci. 69, 1191–1207. clam shells host a multitude of epibionts. Bull. Mar. Sci. 90. http://dx.doi.org/ Pearson, R.G., Munro, J.L., 1991. Growth, mortality and recruitment rates of giant 10.5343/bms.2014.1010. clams, Tridacna gigas and T. derasa, at Michaelmas Reef, central Great Barrier Wahl, M., Mark, O., 1999. The predominantly facultative nature of epibiosis: Reef, Australia. Aust. J. Mar. Fresh. Res. 42, 241–262. experimental and observational evidence. Mar. Ecol. Prog. Ser. 187, 59–66. Penny, S.S., Willan, R.C., 2014. Description of a new species of giant clam Walpole, M.J., Leader-Williams, N., 2002. Tourism and flagship species in (Bivalvia: Tridacnidae) from Ningaloo Reef, Western Australia. Mollus. Res. conservation. Biodivers. Conserv. 11, 543–547. 34, 201–211. Wanders, J.B.W., 1976. The role of benthic algae in the shallow reef of Curacao Perio (de), N.S., Belda, C.A., 1989. Predators and parasites of giant clams (Bivalvia: (Netherlands Antilles) II: primary productivity of the Sargassum beds on the Tridacnidae) in the land-and ocean-based nurseries in Bolinao, Pangasinan. in: north-east coast of submarine plateau. Aquat. Bot. 2, 327–335. Zaragoza, E.C., de Guzman, D.L., Gonzales, E.P. (Eds.), Culture of giant clams Watanabe, T., Suzuki, A., Kawahata, H., Kan, H., Ogawa, S., 2004. A 60-year isotopic (Bivalvia: Tridacnidae). Proceedings of the Symposium on the culture of giant record from a mid-Holocene fossil giant clam (Tridacna gigas) in the Ryukyu clams. Silliman University, Dumaguete City, March 15–17 1988. Philippine Islands: physiological and paleoclimatic implications. Palaeogeogr. Council for Aquatic and Marine Research and Development and Australian Palaeoclimatol. 212, 343–354. Centre for International Agricultural Research (Book Series No. 02/1989), pp. Waters, C.G., Story, R., Costello, M.J., 2013. A methodology for recruiting a giant 75–80. clam, Tridacna maxima, directly to natural substrata: a first step in reversing Perron, F.E., Heslinga, G.A., Fagolimul, J.O., 1985. The gastropod Cymatium functional ? Biol. Conserv. 160, 19–24. muricinum, a predator on juvenile tridacnid clams. Aquaculture 48, 211–221. Weingarten, R.A., 1991. Tridacna – the giant clam. Freshwater Mar. Aquarium 14 Perry, C.T., Edinger, E.N., Kench, P.S., Murphy, G.N., Smithers, S.G., Steneck, R.S., (101–106), 182. Mumby, P.J., 2012. Estimating rates of biologically driven coral reef framework Wells, S., 1996. IUCN 2013. IUCN Red List of Threatened Species. Version 2013.2. production and erosion: a new census-based carbonate budget methodology (accessed 05.05.14). and applications to the reefs of Bonaire. Coral Reefs 31, 853–868. Yamashita, H., Suzuki, G., Hayashibara, T., Koike, K., 2011. Do corals select Pilditch, C.A., Emerson, C.W., Grant, J., 1998. Effect of scallop shells and sediment zooxanthellae by alternative discharge? Mar. Biol. 158, 87–100. grain size on phytoplankton flux to the bed. Cont. Shelf Res. 17, 1869–1885. Yau, A.J.-Y., Fan, T.-Y., 2012. Size-dependent photosynthetic performance in the Pitt, K.A., Welsh, D.T., Condon, R.H., 2009. Influence of jellyfish blooms on carbon, giant clam Tridacna maxima, a mixotrophic marine bivalve. Mar. Biol. 159, 65– nitrogen and phosphorus cycling and production. Hydrobiologia 616, 75. 133–149. Yonge, C.M., 1953. Mantle chambers and water circulation in the Tridacnidae Porter, J.W., Muscatine, L., Dubinsky, Z., Falkowski, P.G., 1984. Primary production (Mollusca). Proc. Zool. Soc. London 123, 551–561. and photoadaptation in light- and shade-adapted colonies of the symbiotic Yonge, C.M., 1955. Adaptation to rock boring in and Lithophaga coral, pistillata. Proc. Roy. Soc. London B 222, 161–180. (Lamellibranchia, ) with a discussion on the evolution of this habit. Rassweiler, A., Rassweiler, T., 2011. Does rapid scavenging hide non-predation Quart. J. Microsc. Sci. 96, 383–410. mortality in coral reef communities? Mar. Freshwater Res. 62, 510–515. Yonge, C.M., 1975. Giant clams. Sci. Am. 232, 96–105. Rees, M., Colquhoun, J., Smith, L.D., Heyward, A.J., 2003. Survey of trochus, Yonge, C.M., 1980. Functional morphology and evolution in the Tridacnidae holothuria, giant clams, and the coral communities of Ashmore, Cartier Reef (Mollusca: Bivalvia: Cardiacea). Rec. Aust. Mus. 33, 735–777.